scispace - formally typeset
Open AccessJournal ArticleDOI

The Structure of the Potassium Channel: Molecular Basis of K+ Conduction and Selectivity

Reads0
Chats0
TLDR
The architecture of the pore establishes the physical principles underlying selective K+ conduction, which promotes ion conduction by exploiting electrostatic repulsive forces to overcome attractive forces between K+ ions and the selectivity filter.
Abstract
The potassium channel from Streptomyces lividans is an integral membrane protein with sequence similarity to all known K+ channels, particularly in the pore region. X-ray analysis with data to 3.2 angstroms reveals that four identical subunits create an inverted teepee, or cone, cradling the selectivity filter of the pore in its outer end. The narrow selectivity filter is only 12 angstroms long, whereas the remainder of the pore is wider and lined with hydrophobic amino acids. A large water-filled cavity and helix dipoles are positioned so as to overcome electrostatic destabilization of an ion in the pore at the center of the bilayer. Main chain carbonyl oxygen atoms from the K+ channel signature sequence line the selectivity filter, which is held open by structural constraints to coordinate K+ ions but not smaller Na+ ions. The selectivity filter contains two K+ ions about 7.5 angstroms apart. This configuration promotes ion conduction by exploiting electrostatic repulsive forces to overcome attractive forces between K+ ions and the selectivity filter. The architecture of the pore establishes the physical principles underlying selective K+ conduction.

read more

Content maybe subject to copyright    Report

The Structure of the Potassium
Channel: Molecular Basis of K
1
Conduction and Selectivity
Declan A. Doyle, Joa˜ o Morais Cabral, Richard A. Pfuetzner,
Anling Kuo, Jacqueline M. Gulbis, Steven L. Cohen,
Brian T. Chait, Roderick MacKinnon*
The potassium channel from Streptomyces lividans is an integral membrane protein with
sequence similarity to all known K
1
channels, particularly in the pore region. X-ray
analysis with data to 3.2 angstroms reveals that four identical subunits create an inverted
teepee, or cone, cradling the selectivity filter of the pore in its outer end. The narrow
selectivity filter is only 12 angstroms long, whereas the remainder of the pore is wider
and lined with hydrophobic amino acids. A large water-filled cavity and helix dipoles are
positioned so as to overcome electrostatic destabilization of an ion in the pore at the
center of the bilayer. Main chain carbonyl oxygen atoms from the K
1
channel signature
sequence line the selectivity filter, which is held open by structural constraints to co-
ordinate K
1
ions but not smallerNa
1
ions. The selectivity filter contains two K
1
ions about
7.5 angstroms apart. This configuration promotes ion conduction by exploiting electro-
static repulsive forces to overcome attractive forces between K
1
ions and the selectivity
filter. The architecture of the pore establishes the physical principles underlying selective
K
1
conduction.
Potassium ions diffuse rapidly across cell
membranes through proteins called K
1
channels. This movement underlies many
fundamental biological processes, includ-
ing electrical signaling in the nervous sys-
tem. Potassium channels use diverse
mechanisms of gating (the processes by
which the pore opens and closes), but they
all exhibit very similar ion permeability
characteristics (1). All K
1
channels show
a selectivity sequence of K
1
' Rb
1
.
Cs
1
, whereas permeability for the smallest
alkali metal ions Na
1
and Li
1
is immea-
surably low. Potassium is at least 10,000
times more permeant than Na
1
, a feature
that is essential to the function of K
1
channels. Potassium channels also share a
constellation of permeability characteris-
tics that is indicative of a multi-ion
conduction mechanism: The flux of ions
in one direction shows high-order cou-
pling to flux in the opposite direction, and
ionic mixtures result in anomalous con-
duction behavior (2). Because of these
properties, K
1
channels are classified
as “long pore channels,” invoking the
notion that multiple ions queue inside a
long, narrow pore in single file. In
addition, the pores of all K
1
channels
can be blocked by tetraethylammonium
(TEA) ions (3).
Molecular cloning and mutagenesis ex-
periments have reinforced the conclusion
that all K
1
channels have essentially the
same pore constitution. Without exception,
all contain a critical amino acid sequence,
the K
1
channel signature sequence. Muta-
tion of these amino acids disrupts the chan-
nel’s ability to discriminate between K
1
and Na
1
ions (4).
Two aspects of ion conduction by K
1
channels have tantalized biophysicists for
the past quarter century. First, what is the
chemical basis of the impressive fidelity
with which the channel distinguishes be-
tween K
1
and Na
1
ions, which are feature-
less spheres of Pauling radius 1.33 Å and
0.95 Å, respectively? Second, how can K
1
channels be so highly selective and at the
same time, apparently paradoxically, exhib-
it a throughput rate approaching the diffu-
sion limit? The 10
4
margin by which K
1
is
selected over Na
1
implies strong energetic
interactions between K
1
ions and the pore.
And yet strong energetic interactions seem
incongruent with throughput rates up to
10
8
ions per second. How can these two
essential features of the K
1
channel pore be
reconciled?
Potassium Channel Architecture
The amino acid sequence of the K
1
chan-
nel from Streptomyces lividans (KcsA K
1
channel) (5) is similar to that of other K
1
channels, including vertebrate and inverte-
brate voltage-dependent K
1
channels, ver-
tebrate inward rectifier and Ca
21
-activated
K
1
channels, K
1
channels from plants and
bacteria, and cyclic nucleotide-gated cation
channels (Fig. 1). On the basis of hydro-
phobicity analysis, there are two closely
related varieties of K
1
channels, those con-
taining two membrane-spanning segments
per subunit and those containing six. In all
cases, the functional K
1
channel protein is
a tetramer (6), typically of four identical
subunits (7). Subunits of the two mem-
brane-spanning variety appear to be short-
ened versions of their larger counterparts, as
if they simply lack the first four membrane-
spanning segments. Although the KcsA K
1
channel is a two membrane-spanning K
1
channel, its amino acid sequence is actually
closer to those of eukaryotic six membrane-
spanning K
1
channels. In particular, its
sequence in the pore region, located be-
tween the membrane-spanning stretches
and containing the K
1
channel signature
sequence, is nearly identical to that found
in the Drosophila (Shaker) and vertebrate
voltage-gated K
1
channels (Fig. 1). In an
accompanying paper, through a study of the
KcsA K
1
channel interaction with eukary-
otic K
1
channel toxins, we confirm that
the KcsA pore structure is indeed very sim-
ilar to that of eukaryotic K
1
channels and
that its structure is maintained when it is
removed from the membrane with deter-
gent (8).
We have determined the KcsA K
1
channel structure from residue position 23
to 119 by x-ray crystallography (Table 1).
The cytoplasmic carboxyl terminus (resi-
dues 126 to 158) was removed in the prep-
aration and the remaining residues were
disordered. The KcsA K
1
channel crystals
are radiation-sensitive and the diffraction
pattern is anisotropic, with reflections ob-
served along the best and worst directions
at 2.5 Å and 3.5 Å Bragg spacings, respec-
tively. By data selection, anisotropy correc-
tion, introduction of heavy atom sites by
site-directed mutagenesis, averaging, and
solvent flattening, an interpretable electron
density map was calculated (Fig. 2, A
through C). This map was without main
chain breaks and showed strong side chain
density (Fig. 2C). The model was refined
with data to 3.2 Å (the data set was 93 %
complete to 3.2 Å with 67% completeness
between 3.3 Å and 3.2 Å), maintaining
highly restrained stereochemistry and keep-
ing tight noncrystallographic symmetry re-
straints. The refinement procedure was
D. A. Doyle, R. A. Pfuetzner, A. Kuo, and R. MacKinnon
are in the Laboratory of Molecular Neurobiology and Bio-
physics and the Howard Hughes Medical Institute, Rock-
efeller University, 1230 York Avenue, New York, NY
10021, USA. J. M. Cabral and J. M. Gulbis are in the
Laboratory of Molecular Neurobiology and Biophysics,
Rockefeller University, 1230 York Avenue, New York, NY
10021, USA. S. L. Cohen and B. T. Chait are in the
Laboratory of Mass Spectrometry and Gaseous Ion
Chemistry, Rockefeller University, 1230 York Avenue,
New York, NY 10021, USA.
*To whom correspondence should be addressed. E-mail:
mackinn@rockvax.rockefeller.edu
RESEARCH ARTICLES
www.sciencemag.org
z
SCIENCE
z
VOL. 280
z
3 APRIL 1998 69

monitored by minimizing the value R-free
(29.0%) and its separation from R-crystal-
lographic (28.0%). The presence of four
molecules (subunits) in the asymmetric unit
of the crystal provides a very significant
enhancement of the accuracy of the crys-
tallographic analysis; first, by enabling av-
eraging of the electron density over four
crystallographically independent regions of
the multiple isomorphous replacement
(MIR) map, and second, by providing a
powerful set of constraints on the atomic
model during refinement (9).
The K
1
channel is a tetramer with four-
fold symmetry about a central pore (Fig. 3,
A and B). Like several other membrane
proteins, it has two layers of aromatic amino
acids positioned to extend into the lipid
bilayer, presumably near the membrane-wa-
ter interfaces (Fig. 3C) (10). Each subunit
has two transmembrane a-helices connect-
ed by the roughly 30 amino acid pore re-
gion, which consists of the turret, pore he-
lix, and selectivity filter (Fig. 3, A and B).
A subunit is inserted into the tetramer such
that one transmembrane helix (inner helix)
faces the central pore while the other (outer
helix) faces the lipid membrane. The inner
helices are tilted with respect to the mem-
brane normal by about 25° and are slightly
kinked, so that the subunits open like the
petals of a flower facing the outside of the
cell. The open petals house the structure
formed by the pore region near the extra-
cellular surface of the membrane. This re-
gion contains the K
1
channel signature
sequence, which forms the selectivity filter
(4). The essential features of subunit pack-
ing can be appreciated by viewing the rela-
tion between the four inner helices and the
four pore helices (Fig. 3D). The four inner
helices pack against each other as a bundle
near the intracellular aspect of the mem-
brane, giving the appearance of an inverted
teepee. The pore helices are slotted in be-
tween the poles of the teepee and are di-
rected, with an amino-to-carboxyl sense,
toward a point near the center of the chan-
nel (Fig. 3, A, B, and D). This pore helix
arrangement provides many of the intersub-
unit contacts that hold the tetramer togeth-
er and, as discussed below, is also critical in
the operation of the ion conduction pore.
Sequence conservation among K
1
chan-
nels (including ones with two and six mem-
brane-spanning segments), as well as cyclic
nucleotide-gated cation channels, is stron-
gest for the amino acids corresponding to
the pore region and the inner helix. Even
Na
1
and Ca
21
channels show distant relat-
edness over these segments. The teepee ar-
chitecture of the K
1
channel pore likely
will be a general feature of all of these
cation channels, with four inner helices
arranged like the poles of a teepee, four pore
helices, and a selectivity filter—tuned to
select the appropriate cation—located close
to the extracellular surface.
This structure of the KcsA K
1
channel
is in excellent agreement with results from
functional and mutagenesis studies on Shak-
er and other eukaryotic K
1
channels (Fig.
4). The pore-region of K
1
channels was
first defined with pore-blocking scorpion
toxins (11). These inhibitors interact with
amino acids (Fig. 4, white) comprising the
broad extracellular-facing entryway to the
pore (12). The impermeant organic cation
TEA blocks K
1
channels from both sides of
the membrane at distinct sites (13). Amino
acids interacting with externally and inter-
nally applied TEA are located just external
to (Fig. 4, yellow) and internal to (Fig. 4,
mustard) the structure formed by the signa-
ture sequence amino acids (14, 15). Alter-
ation of the signature sequence amino acids
(Fig. 4, red main chain atoms) disrupts K
1
selectivity (4). Amino acids close to the
intracellular opening on the Shaker K
1
channel map to the inner helix on the
KcsA K
1
channel (16). Interestingly, expo-
sure to the cytoplasm of the region above
the inner helix bundle (Fig. 4, pink side
chains) requires an open voltage-dependent
gate, whereas the region at or below the
bundle (Fig. 4, green side chains) is exposed
whether or not the gate was open. The
correlation between the transition zone for
gate-dependent exposure to the cytoplasm
in the Shaker K
1
channel and the inner
helix bundle in this structure has im-
plications for mechanisms of gating in K
1
channels.
General Properties of the Ion
Conduction Pore
As might have been anticipated for a cation
channel, both the intracellular and extra-
cellular entryways are negatively charged by
acidic amino acids (Fig. 5A, red), an effect
that would raise the local concentration of
cations while lowering the concentration of
anions. The overall length of the pore is 45
Å, and its diameter varies along its distance
(Fig. 5B). From inside the cell (bottom) the
pore begins as a tunnel 18 Å in length (the
internal pore) and then opens into a wide
cavity (;10 Å across) near the middle of
the membrane. A K
1
ion could move
throughout the internal pore and cavity and
still remain mostly hydrated. In contrast,
the selectivity filter separating the cavity
kcsa TYPRALWWSVETATTVGYGDLY..PVTLWGRLVAVVVMVAGITSFGLVTAALATWFVGRE
kch SLMTAFYFSIETMSTVGYGDIV..PVSESARLFTISVIISGITVFATSMTSIFGPLIRGG
clost SLGNALWWSFVTITTVGYGDIS..PSTPFGRVIASILMLIGIGFLSMLTGTISTFFISKK
Shaker SIPDAFWWAVVTMTTVGYGDMT..PVGFWGKIVGSLCVIAGVLTIALPVPVIVSNFNYFY
hKv1.1 SIPDAFWWAVVSMTTVGYGDMY..PVTIGGKIVGSLCAIAGVLTIALPVPVIVSNFNYFY
hDRK SIPASFWWATITMTTVGYGDIY..PKTLLGKIVGGLCCIAGVLVIALPIPIIVNNFSEFY
Parame QYLHSLYWSIITMTTIGYGDIT..PQNLRERVFAVGMALSAVGVFGYSIGNINSIYAEWS
Celegans SIPLGLWWAICTMTTVGYGDMT..PHTSFGRLVGSLCAVMGVLTIALPVPVIVSNFAMFY
mSlo TYWECVYLLMVTMSTVGYGDVY..AKTTLGRLFMVFFILGGLAMFASYVPEIIELIGNRK
cal_act NFLGAMWLISITFLSIGYGDMV..PHTYCGKGVCLLTGIMGAGCTALVVAVVARKLELTK
AKT1 RYVTSMYWSITTLTTVGYGDLH..PVNTKEMIFDIFYMLFNLGLTAYLIGNMTNLVVHGT
herg KYVTALYFTFSSLTSVGFGNVS..PNTNSEKIFSICVMLIGSLMYASIFGNVSAIIQRLY
romk GMTSAFLFSLETQVTIGYGFRFVTEQCATAIFLLIFQSILGVIINSFMCGAILAKISRPK
hgirk GFVSAFLFSIETETTIGYGYRVITDKCPEGIILLLIQSVLGSIVNAFMVGCMFVKISQPK
olCNG EYIYCLYWSTLTLTTIG..ETPP.PVKDEEYLFVIFDFLIGVLIFATIVGNVGSMISNMN
rodCNG KYVYSLYWSTLTLTTIG..ETPP.PVRDSEYVFVVVDFLIGVLIFATIVGNIGSMISNMN
NC
OUTER HELIX
INNER HELIX
PORE
HELIX
PORE REGION
61 70 80 90 100 110
.
.
.
.
.
.
.
.
.
.
.
.
Fig. 1. Sequence alignment of selected K
1
channels and cyclic nucleotide-gated channels. The
numbering and secondary structural elements for the Streptomyces lividans K
1
channel (KcsA) is given
above the sequences. Selectivity filter, red; lining of the cavity and inner pore, blue; residues in which the
nature of the side chain is preserved (.50% similarity), grey. The sequences are: KcsA, Streptomyces
lividans, accession number (acc) PIR S60172; kch, Escherichia coli, acc GenBank U24203; clost,
Clostridium acetobutylicum (Genome Therapeutics Corp.); Shaker, Drosophila melanogaster, acc PIR
S00479; hKv1.1, Homo sapiens, acc Swissprot Q09470; hDRK, H. sapiens, acc PIR S31761;
Parame, Paramecium tetraaurelia, acc GenBank U19908; Celegans, Caenorhabditis elegans, acc
GenBank AF005246; mSlo, Mus musculus, acc PIR A48206; cal_act, H. sapiens, acc GenBank
AF031815; AKT1, Arabidopsis thaliana, acc PIR S62694; herg, H. sapiens, acc PIR I38465; romk,
Rattus norvegicus, acc Swissprot P35560; hgirk, H. sapiens, acc GenBank S78684; olCNG, H. sapi-
ens, acc Swissprot Q16280; rodCNG, H. sapiens, acc PIR A42161. The last two sequences, separate
from the rest, are from cyclic nucleotide–gated channels, which are not K
1
selective.
SCIENCE
z
VOL. 280
z
3 APRIL 1998
z
www.sciencemag.org70

from the extracellular solution is so narrow
that a K
1
ion would have to shed its hy-
drating waters to enter. The chemical com-
position of the wall lining the internal pore
and cavity is predominantly hydrophobic
(Fig. 5A, yellow). The selectivity filter, on
the other hand, is lined exclusively by polar
main chain atoms belonging to the signa-
ture sequence amino acids. The distinct
mechanisms operating in the cavity and
internal pore versus the selectivity filter will
be discussed below, but first we introduce
the determination of K
1
ion positions in
the pore.
Potassium channels exclude the smaller
alkali metal cations Li
1
(radius 0.60 Å) and
Na
1
(0.95 Å) but allow permeation of the
larger members of the series Rb
1
(1.48 Å)
and Cs
1
(1.69 Å). In fact, Rb
1
is nearly a
perfect K
1
(1.33 Å) analog because its size
and permeability characteristics are very
similar to those of K
1
. Because they are
more electron dense than K
1
,Rb
1
and Cs
1
allow visualization of the locations of per-
meant ions in the pore. By difference elec-
tron density maps calculated with data from
crystals transferred into Rb
1
-containing
(Fig. 6a) or Cs
1
-containing (Fig. 6b) solu-
tions, multiple ions are well defined in the
pore. The selectivity filter contains two
ions (inner and outer ions) located at op-
posite ends, about 7.5 Å apart (center to
center). In the Rb
1
difference map, there
actually are two partially separated peaks at
the inner aspect of the selectivity filter.
These peaks are too close to each other (2.6
Å) to represent two simultaneously occu-
pied ion binding sites. We suspect that they
represent a single ion (on average) in rapid
equilibrium between adjacent sites. The
single inner ion peak in the Cs
1
difference
map undoubtedly reflects the lower resolu-
tion at which the map was calculated (to 5
Å for Cs
1
versus 4.0 Å for Rb
1
), because
the Rb
1
difference map, when calculated at
the same lower resolution, also shows only a
single peak at the Cs
1
position. The Rb
1
positions correspond to strong peaks (pre-
A
B
C
Fig. 2. Experimental electron density map. Ste-
reoviews of the experimental electron-density
map contoured at 1 s covering nearly an entire
subunit (removed from the tetramer) of the final
model. The map was calculated at 3.2 Å resolu-
tion with the following Fourier coefficients: native-
sharpened amplitudes and MIR solvent flattened
averaged phases. (A) Foreground: map showing
inner helix, loop structures and selectivity filter;
background: the pore helix and outer helix. CPK
spheres show positions of mercury atoms used as
residue markers (from the top, marked residues
are Leu
86
, Leu
90
and Val
93
). (B) Alternative view.
Foreground: pore helix and part of outer helix;
background: selectivity filter and turret. CPK
sphere marks position of Ala
42
.(C) Close up view
of electron density.
RESEARCH ARTICLES
www.sciencemag.org
z
SCIENCE
z
VOL. 280
z
3 APRIL 1998 71

sumably, K
1
ions) in a high contour native
electron density map (not shown). Thus,
the selectivity filter contains two K
1
ions.
A third weaker peak is located below the
selectivity filter at the center of the large
cavity in the Rb
1
difference map (Fig. 6a,
cavity ion) and in the Cs
1
difference map
at a lower contour (not shown). Electron
density at the cavity center is prominent in
MIR maps, even prior to averaging (Fig. 6c,
lower diffuse peak). The difference electron
density maps show this to be related to the
presence of one or more poorly localized
cations situated at least 4 Å away from
the closest protein groups.
The Cavity and Internal Pore
Why is there a 10 Å diameter cavity in the
center of the channel with an ion in it
(Fig. 5B and Fig. 6)? Electrostatic calcula-
tions show that when an ion is moved
along a narrow pore through a membrane,
it must cross an energy barrier that is
maximum at the membrane center (17).
The electrostatic field emanating from a
cation polarizes its environment, bringing
the negative ends of dipoles closer to it
and thereby stabilizing it. At the bilayer
center, the polarizability of the surround-
Table 1. Summary of data collection and refinement statistics. Crystals
(space group C2: a 5 128.8 Å, b 5 68.9 Å, c 5 112.0 Å, b5124.6°) were
flash-frozen by being transferred directly from the crystal mother liquor to a
stream of boiled-off nitrogen (24). Because crystals of the mutant L90C
diffracted significantly better than wild-type protein crystals, the former were
used for native data collection. Data were collected from multiple crystals,
and six sets were selected and merged to form the native data set used for
structure determination. Mercury derivatives were obtained by direct addition
of methyl mercury to the crystallization solution of cysteine mutant crystals.
MALDI-TOF mass spectrometry confirmed 60 to 90% derivatization of crys-
tals prior to data collection. All data were collected at Cornell High Energy
Synchrotron Source (CHESS), station A1, with the Princeton 2K CCD (25).
Data were processed with DENZO and SCALEPACK (26) and the CCP4
package (27). Heavy atom positions were determined with SHELX-97 (28)
and cross-difference Fourier analysis. These positions confirmed the fourfold
noncrystallographic symmetry observed in the self-rotation Patterson func-
tion and allowed the determination of initial orientation matrices. An initial
model (90% complete) was built into a solvent flattened (64% solvent con-
tent), fourfold averaged electron density map with the program O (29). The
tracing of the model was facilitated by the use of the mercury atom positions
as residue markers. L86C was used solely for this purpose. After torsional
refinement (with strict fourfold noncrystallographic symmetry constraints)
with X-PLOR 3.851 (30), this model was used in the anisotropic scaling
[sharpening (31)] of the native data with X-PLOR. The structure factor sigma
values were also rescaled appropriately, and the corrected data were used
for all subsequent procedures. Fourfold averaging, solvent flattening, and
phase extension were applied in DM (32), resulting in a marked improvement
of the electron density that allowed correction of the model and the building
of additional residues. Refinement consisted of rounds of positional (in the
initial stages phase information was also included as a restraint) and grouped
B-factor refinement in X-PLOR. Fourfold noncrystallographic symmetry was
highly restrained with the force constant for positional restraints set as 1000
kcal/mol/Å
2
. The diffuse ion cloud described in the text was initially modeled
as one or more K
1
ions and several water molecules; however, the results
were unsatisfactory. Therefore, this and other strong unmodelled density
present in solvent-flattened maps (no averaging included) was Fourier back-
transformed, scaled, and included in the refinement procedure as partial
structure factors. The final model includes amino acids 23 to 119 of each
chain. The following residues were truncated: Arg
27
to Cb, Ile
60
to Cg, Arg
64
to Cb, Glu
71
to Cb, and Arg
117
to N«. The stereochemistry is strongly
restrained, with no outliers on the Ramachandran plot. The high B-factor
values reflect the intensity decay of the data beyond 4 Å.
Data collection and phasing
Data set
Resolution
(Å)
Redundancy
Completeness
overall/outer
(%)
R
merge
*
Phasing
power
R-Cullis
L90C-a 15.0–3.7 3.5 91.3/93.3 0.071 1.61 0.70
L90C-b 15.0–3.7 7.0 91.5/94.1 0.083 1.87 0.50
V93C 15.0–3.7 4.1 98.3/99.1 0.075 1.35 0.63
A32C 15.04.0 2.3 84.1/83.8 0.076 1.45 0.66
A29C 15.0–5.0 2.7 73.9/74.0 0.063 1.03 0.85
A42C 15.06.5 2.0 90.7/90.3 0.057 0.97 0.81
L86C 30.06.0 2.3 58.7/58.9 0.057
Native 30.0–3.2 6.1 93.3% 0.086 15.8 75
Outer Shell 3.3–3.2 2.3 66.6% 0.286 3.9 50
Anisotropic correction
Before sharpening
After sharpening
Average F.O.Mi
(30.0–3.2 Å)
Average F.O.Mi
(3.4–3.2 Å)
0.76
0.83
0.55
0.64
Refinement Root-mean-square deviation of
Resolution: 10.0–3.2 Å Bond angles: 1.096°
R-cryst.§: 28.0% Bond lengths: 0.005 Å
R-free§: 29.0% Ncs related atoms: 0.006 Å
No. of reflections with F/sF . 2: 12054 B-factor for ncs related atoms: 10 Å
2
No. of protein atoms: 710 per subunit B-factor for non-bonded atoms: 36 Å
2
No. of ligand atoms: 1 water, 3 K
1
ions
Mean B-factor for main-chain atoms: 90 Å
2
Mean B-factor for side-chain atoms: 110 Å
2
*R
merge
5SSI2Ij/SI. Phasing power 5^F
h
&/^E&. R-Cullis 5SiF
ph
6 F
p
2F
hc
i/SF
ph
6 F
p
, only for centric data. §R-cryst. 5SF
p
2F
p(calc)
/SF
p
, R-free the
same as R-cryst, but calculated on 10% of data selected in thin resolution shells and excluded from refinement. iFigure of merit. In both cases, fourfold averaging and
solvent flattening were applied; Ij is the observed intensity, I is the average intensity, F
h
is the root-mean-square heavy-atom structure factor, E is the lack of closure error, F
ph
is the
structure factor for the derivative, F
p
is the structure factor for the native, F
hc
is the calculated structure factor for the heavy atom, and F
p(calc)
is the calculated native structure factor.
I/sI % of measured
data with I/sI . 2
SCIENCE
z
VOL. 280
z
3 APRIL 1998
z
www.sciencemag.org72

ing medium is minimal and therefore the
energy of the cation is highest. Thus, sim-
ple electrostatic considerations allow us to
understand the functional significance of
the cavity and its strategic location. The
cavity overcomes the electrostatic desta-
bilization resulting from the low dielectric
bilayer by simply surrounding an ion with
polarizable water. A second feature of the
K
1
channel structure also stabilizes a cat-
ion at the bilayer center. The four pore
helices point directly at the center of the
cavity (Fig. 3, A, B, and D). The amino to
carboxyl orientation of these helices will
impose a negative electrostatic (cation at-
tractive) potential via the helix dipole
effect (18). The ends of the helices are
rather far (;8 Å) from the cavity center,
but all four contribute to the effect.
Therefore, two properties of the structure,
the aqueous cavity and the oriented heli-
ces, help to solve a fundamental physical
problem in biology—how to lower the
electrostatic barrier facing a cation cross-
ing a lipid bilayer. Thus, the diffuse elec-
tron density in the cavity center (Fig. 6C)
likely reflects a hydrated cation cloud
rather than an ion binding site (Fig. 7).
Alternatively, the channel could have
overcome the destabilizing electrostatic
effects of the bilayer center by lining the
entire pore with a polarizable surface, put-
ting ion binding sites along its entire
length. But the structure shows that, with
the exception of the selectivity filter, the
pore lining is mainly hydrophobic, a gen-
eral property of K
1
channels (Fig. 1). This
conclusion was anticipated by the land-
mark experiments of Armstrong, which
showed that hydrophobic cations bind in
A
B
D
Fig. 3. Views of the tetramer. (A) Stereoview of a
ribbon representation illustrating the three-dimen-
sional fold of the KcsA tetramer viewed from the
extracellular side. The four subunits are distin-
guished by color. (B) Stereoview from another
perspective, perpendicular to that in (A). (C) Rib-
bon representation of the tetramer as an integral-
membrane protein. Aromatic amino acids on the
membrane-facing surface are displayed in black.
(D) Inverted teepee architecture of the tetramer.
These diagrams were prepared with MOLSCRIPT
and RASTER-3D (33).
RESEARCH ARTICLES
www.sciencemag.org
z
SCIENCE
z
VOL. 280
z
3 APRIL 1998 73

Figures
Citations
More filters
Journal Article

The glutamate receptor ion channels

TL;DR: The cloning of cDNAs encoding glutamate receptor subunits, which occurred mainly between 1989 and 1992, stimulated the development of ionotropic glutamate receptors in the brain.
Journal ArticleDOI

Glutamate Receptor Ion Channels: Structure, Regulation, and Function

TL;DR: This review discusses International Union of Basic and Clinical Pharmacology glutamate receptor nomenclature, structure, assembly, accessory subunits, interacting proteins, gene expression and translation, post-translational modifications, agonist and antagonist pharmacology, allosteric modulation, mechanisms of gating and permeation, roles in normal physiological function, as well as the potential therapeutic use of pharmacological agents acting at glutamate receptors.
Journal ArticleDOI

Crystal Structure of a Mammalian Voltage-Dependent Shaker Family K + Channel

TL;DR: In this paper, the authors reported the crystal structure of a mammalian voltage-dependent potassium ion (K+) channel, Kv1.2, which is a member of the Shaker K+ channel family.
Journal ArticleDOI

Dynamic personalities of proteins.

TL;DR: The dream is to 'watch' proteins in action in real time at atomic resolution, which requires addition of a fourth dimension, time, to structural biology so that the positions in space and time of all atoms in a protein can be described in detail.
Journal ArticleDOI

From Ionic Currents to Molecular Mechanisms: The Structure and Function of Voltage-Gated Sodium Channels

TL;DR: Together, these studies showed that the mechanisms of sodium channel function and regulation, purified sodium channel protein contained the essential and gives a perspective for future research on the ex-elements for ion conduction and voltage-dependent panding family of Sodium channel proteins.
References
More filters
Journal ArticleDOI

The CCP4 suite: programs for protein crystallography

TL;DR: The CCP4 (Collaborative Computational Project, number 4) program suite is a collection of programs and associated data and subroutine libraries which can be used for macromolecular structure determination by X-ray crystallography.
Journal ArticleDOI

MOLSCRIPT: a program to produce both detailed and schematic plots of protein structures

TL;DR: The MOLSCRIPT program as discussed by the authors produces plots of protein structures using several different kinds of representations, including simple wire models, ball-and-stick models, CPK models and text labels.
Journal ArticleDOI

Improved methods for building protein models in electron density maps and the location of errors in these models.

TL;DR: In this paper, the authors describe strategies and tools that help to alleviate this problem and simplify the model-building process, quantify the goodness of fit of the model on a per-residue basis and locate possible errors in peptide and side-chain conformations.
Book ChapterDOI

The ionic channels in excitable membranes.

TL;DR: Comparisons can now be made between the kinetics of the ionic conductances as described by Hodgkin & Huxley, and the steady-state distribution and kinetic changes of the charged controlling particles, which should lead to useful conclusions about the intramolecular organization of the sodium channels and the conformational changes that take place under the influence of the electric field.
Journal ArticleDOI

HOLE: a program for the analysis of the pore dimensions of ion channel structural models.

TL;DR: A method (HOLE) that allows the analysis of the dimensions of the pore running through a structural model of an ion channel is presented and can be used to predict the conductance of channels using a simple empirically corrected ohmic model.
Related Papers (5)
Frequently Asked Questions (18)
Q1. What contributions have the authors mentioned in the paper "The structure of the potassium channel: molecular basis of k conduction and selectivity" ?

In this paper, the authors used X-ray analysis with data to 3.2 angstroms reveals that four identical subunits create an inverted teepee, or cone, cradling the selectivity filter of the pore in its outer end. 

Fourfold averaging, solvent flattening, and phase extension were applied in DM (32), resulting in a marked improvement of the electron density that allowed correction of the model and the building of additional residues. 

The side chains colored green, when mutated to cysteine, are modified by cysteine-reactive agents whether or not the channel gate is open, whereas those colored pink react only when the channel is open (16). 

On the basis of functional measurements, the same concept of destabilization by multiple ion occupancy has been proposed for Ca21 channels (22) and for K1 channels (23) and may be a general property of all selective ion channels. 

A single water molecule (the only one modeled in the structure) located between the two K1 ions in the selectivity filter was justified by the presence of a strong electron density peak in the experimental map, which was never associated with an ion peak in the difference Fourier maps (19). 

The amino to carboxyl orientation of these helices will impose a negative electrostatic (cation attractive) potential via the helix dipole effect (18). 

Although the KcsA K1 channel is a two membrane-spanning K1 channel, its amino acid sequence is actually closer to those of eukaryotic six membranespanning K1 channels. 

The mustard-colored side chain at the base of the selectivity filter affects TEA binding from the intracellular solution [the internal TEA site (15)]. 

The authors also used K1 ion positions defined by difference Fourier analysis (Figs. 6 and 8A, yellow density) and their knowledge of alkali metal cation coordination in small molecules. 

The ions are located at opposite ends of the selectivity filter, separated by about 7.5 Å, roughly the average distance between K1 ions in a 4 M KCl solution, and in the selectivity filter there are no intervening Cl2 anions to balance the charge. 

General Properties of the Ion Conduction PoreAs might have been anticipated for a cation channel, both the intracellular and extracellular entryways are negatively charged by acidic amino acids (Fig. 5A, red), an effect that would raise the local concentration of cations while lowering the concentration of anions. 

when a second ion enters, the attractive force between a K1 ion and the selectivity filter becomes perfectly balanced by the repulsive force between ions, and this is what allows conduction to occur. 

Studies in the rabbit have shown that the cerebellum is essential for both the acquisition and retention of delay classical conditioning (4) and that no other forebrain structure, including the hippocampus, is required (5). 

the channel could have overcome the destabilizing electrostatic effects of the bilayer center by lining the entire pore with a polarizable surface, putting ion binding sites along its entire length. 

The four inner helices pack against each other as a bundle near the intracellular aspect of the membrane, giving the appearance of an inverted teepee. 

this and other strong unmodelled density present in solvent-flattened maps (no averaging included) was Fourier backtransformed, scaled, and included in the refinement procedure as partial structure factors. 

The structure reveals that the selectivity filter is held open as if to prevent it from accommodating a Na1ion with its smaller radius. 

Because of these properties, K1 channels are classified as “long pore channels,” invoking the notion that multiple ions queue inside a long, narrow pore in single file.